Next Article in Journal / Special Issue
In Vivo Gastroprotective and Antidepressant Effects of Iridoids, Verbascoside and Tenuifloroside from Castilleja tenuiflora Benth
Previous Article in Journal
Apigenin Inhibits IL-31 Cytokine in Human Mast Cell and Mouse Skin Tissues
Previous Article in Special Issue
Triterpene Acid and Phenolics from Ancient Apples of Friuli Venezia Giulia as Nutraceutical Ingredients: LC-MS Study and In Vitro Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polyacetylenes from the Roots of Swietenia macrophylla King

1
Key Laboratory of Biology and Genetic Resources of Tropical Crops, Ministry of Agriculture, Institute of Tropical Bioscience and Biotechnology, Chinese Academy of Tropical Agricultural Sciences, Haikou 571101, China
2
Institute of Tropical Agriculture and Forestry, Hainan University, Haikou 570228, China
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(7), 1291; https://doi.org/10.3390/molecules24071291
Submission received: 22 February 2019 / Revised: 20 March 2019 / Accepted: 22 March 2019 / Published: 2 April 2019
(This article belongs to the Special Issue Biological Activities of Plant Secondary Metabolites)

Abstract

:
A phytochemical investigation of the roots of Swietenia macrophylla led to the isolation of seven polyacetylenes, including five new compounds (15) and two known ones (67). Their structures were elucidated by extensive spectroscopic analysis and detailed comparison with reported data. All the isolates were tested for their cytotoxicity against the human hepatocellular carcinoma cell line BEL-7402, human myeloid leukemia cell line K562, and human gastric carcinoma cell line SGC-7901. Compounds 1 and 6 showed moderate cytotoxicity against the above three human cancer cell lines with IC50 values ranging from 14.3 to 45.4 μM. Compound 4 displayed cytotoxicity against the K562 and SGC-7901 cancer cell lines with IC50 values of 26.2 ± 0.4 and 21.9 ± 0.3 μM, respectively.

Graphical Abstract

1. Introduction

Natural polyacetylenes, structures featuring two or more triple bonds [1], are mainly found in plants belonging to the Araliaceae [2,3,4,5,6,7], Asteraceae [8,9,10], Umbelliferae [11,12,13], Santalaceae [14], Pittosporaceae [15], and Oleaceae [16] families. However, polyacetylenes are uncommon in the Meliaceae family, with only six such compounds having been found in total in this family, one of which is from Swietenia mahagoni [17], three from Toona ciliate [18], and two from the stem bark of Khaya ivorensis A. Chev [19,20]. Naturally occurring polyynes are classified into four types: acyclic C18–C14 acetylene compounds; acyclic C13–C8 acetylene compounds; compounds with an allene substructure; and aromatic and heterocyclic acetylene compounds [1]. Diverse polyacetylene structures exhibit a series of bioactivities, including cytotoxicity [8,10,11,21] and antimicrobial [14,22,23], antiviral [24], and enzyme-inhibitory [6,25] activities.
Swietenia macrophylla, a perennial deciduous timber tree that reaches a height of up to 50 m [26], is native to Central and South America [27] and widely distributed in West India, Malaysia, and southern China [28,29]. Antecedent chemical investigations on S. macrophylla have focused mostly on the aboveground parts and their bioactive limonoids [30]. It is necessary to expand the scope of research on S. macrophylla and discover or develop additional biologically active constituents of this plant genus [31]. Our recent study on the roots of S. macrophylla led to the isolation of a series of xanthones, limonoids, and other chemical components [32,33]. As a continuation of our studies on the biologically active agents from this plant, five new and two known acyclic C18–C14 polyacetylenes have been further isolated here, and their cytotoxic activities against the human hepatocellular carcinoma cell line BEL-7402, human myeloid leukemia cell line K562, and human gastric carcinoma cell line SGC-7901 were investigated. In this paper, the isolation, structural elucidation, and cytotoxicity of these compounds are reported as follows.

2. Results and Discussion

The chemical examination of the ethyl acetate (EtOAc) extract from the roots of S. macrophylla resulted in the isolation and identification of five new polyacetylenes (Figure 1), respectively named heptadeca-9-ene-4,6-diyne-3,11-diol (1), (E)-heptadeca-8-ene-4,6-diyne-3,10,11-triol (2), 10-methoxyheptadeca-4,6-diyne-3,9-diol (3), tetradeca-1,3-diyne-6,7,8-triol (4) and 6,7,8,9-tetraacetoxytetradeca-1,3-diyne (5), together with two known compounds, which were identified as α-hexy-3-(6-hydroxy-2,4-ocadiynyl)oxiranemethanol (6) [17] and (3R,8E,10S)-heptadec-8-ene-4,6-diyne-3,10-diol (7) [18] by comparing their experimental spectroscopic data with the reported data in the literature. HRESIMS and NMR spectra for compounds 15 are shown in the Supplementary Materials.
Compound 1 was obtained as a yellow oil with a positive optical rotation [ α ] D 25 +10 (c 0.1, CH3OH) and has a molecular formula of C17H26O2, as evidenced by the HRESIMS peak at m/z 285.1830 [M + Na]+ (calcd 285.1831 for C17H26NaO2), requiring five unsaturation degrees. The UV spectrum displayed typical absorption bands for a conjugated ene-yne-yne chromophore at λmax 280 and 270 nm [11], and the IR spectrum showed the OH group (3435 cm−1), triple-bond (2234 cm−1), and olefinic double-bond (1639 cm−1) absorptions. The 1H-NMR data (Table 1) showed the presence of two olefinic protons at δH 5.50 (1H, m, overlapped, H-9) and 5.48 (1H, m, overlapped, H-10); two oxymethine protons, which appeared at δH 4.34 (1H, t, J = 6.4 Hz, H-3) and 4.39 (1H, q, J = 6.7 Hz, H-11); two triplet methyl groups at δH 0.99 (3H, t, J = 7.4 Hz, H-1) and 0.87 (3H, t, J = 6.8 Hz, H-17); and fourteen aliphatic methylene protons at δH 1.72 (2H, m, H-2), 3.09 (2H, dd, J = 5.0, 10.2 Hz, H-8), 1.60 (1H, m, H-12a), 1.44 (1H, m, H-12b), and 1.27−1.28 (8H, m, overlapped, H-13−16). A detailed analysis of the 13C-NMR and DEPT spectra (Table 1) of compound 1 showed signals for two methyl groups at δC 9.5 (C-1) and 14.2 (C-17); two O-bearing methine carbons at δC 64.1 (C-3) and 67.8 (C-11); two olefinic carbon resonances at δC 135.3 (C-9) and 124.4 (C-10); and seven methylene carbons with chemical shifts ranging from 18.3 to 37.4 ppm. The abovementioned signals accounted for one degree of unsaturation, and thus, the remaining required the existence of two additional triple bonds in the molecule, which consisted of four quaternary carbons δC 77.3 (C-4), 69.7 (C-5), 65.0 (C-6), and 78.6 (C-7) corresponding to the observed substitutions in the UV and IR spectra. By comparison, these data were relatively close to those of panaxjapyne A [6], except for the additional oxymethine at δC 67.8 (C-11), which was clearly proven by the 1H-1H COSY correlations (Figure 2) of H-11 with H-10 and H-12 and HMBC correlations (Figure 2) from H-11 to C-9, C-10, C-12, and C-13. Accordingly, compound 1 was elucidated as heptadeca-9-ene-4,6-diyne-3,11-diol.
Compound 2 was separated as a yellow oil with a negative optical rotation [ α ] D 25 −80 (c 0.1, CHCl3), and its molecular formula was determined to be C17H26O3 by the HRESIMS peak at m/z 301.1776 [M + Na]+ (calcd C17H26NaO3 for 301.1780). The 1H-NMR (Table 1) spectrum showed a pair of trans olefinic protons at δH 6.30 (1H, dd, J = 5.8, 15.9 Hz, H-9) and 5.86 (1H, d, J = 15.9 Hz, H-8); three oxymethine protons at δH 4.41 (1H, t, J = 6.5 Hz, H-3), 4.00 (1H, td, J = 5.8, 1.4 Hz, H-10), and 3.47 (1H, m, H-11); and two methyl groups at δH 1.01 (3H, t, J = 7.4 Hz, H-1) and 0.87 (3H, t, J = 6.9 Hz, H-17). A detailed analysis of the NMR spectroscopic spectra of compound 2 revealed that it is similar to panaxjapyne B [6] except for the additional oxymethine group in compound 2. The 1H-1H COSY correlations (Figure 2) of H-8/H-9/H-10/H-11/H-12 and the HMBC correlations (Figure 2) from H-10 to C-11 and C-12 and from H-11 to C-9 and C-13 evidenced the location of oxymethine at C-11, which accounted for the molecular weight difference of 16 amu observed between the two compounds. Thus, the structure of compound 2 was elucidated as (E)-heptadeca-8-ene-4,6-diyne-3,10,11-triol. The relative configuration of C-10 and C-11 was determined by the 3JH, H value and the ROESY interactions. The small 3JH-10/H-11 value (1.4 Hz) was indicative of a gauche relationship of H-10 and H-11 [34,35]. The ROESY correlations (Figure 3) of H-9/H-12/H-10/H-11 suggested that the relative configurations of C-10 and C-11 were R* and S*, respectively.
Compound 3 was separated as a yellow oil with a positive optical rotation [ α ] D 25 +80 (c 0.1, CHCl3). The molecular formula of compound 3 was established as C18H30O3 from the HRESIMS peak at m/z 317.2089 [M + Na]+ (calcd. C18H30NaO3 for 317.2093). The 1H and 13C-NMR spectroscopic data (Table 1) of compound 3 were comparable to those of oploxyne B [5], suggesting that an oxidized methine of oploxyne B was replaced by the methylene (δH 2.51, 2.57/δC 24.8) of compound 3. Furthermore, the 16 amu of reduced molecular weight compared with oploxyne B indicated that the molecule formulas of the two compounds only differ by one oxygenium. The 1H-1H COSY correlations (Figure 2) of H-8/H-9/H-10/H-11 and the HMBC correlations (Figure 2) from H-8 to C-6, C-7, C-9, and C-10 implied that C-8 is the methylene carbon for compound 3. Finally, the structure of compound 3 was elucidated as 10-methoxyheptadeca-4,6-diyne-3,9-diol.
Compound 4 was obtained as a yellow oil with a negative optical rotation [ α ] D 25 −80 (c 0.1, CHCl3). The molecular formula was determined as C14H22O3 by the HRESIMS peak at m/z 261.1476 [M + Na]+ (calcd for C14H22NaO3, 261.1467). The 1H and 13C-NMR data (Table 2) showed three O-bearing methine groups at δH 4.13 (1H, br t, J = 6.5 Hz, H-6)/δC (69.6), 3.82 (1H, m, H-8)/δC (75.1), and 3.49 (1H, m, H-7)/δC (73.1); one triplet methyl group at δH 0.89 (3H, t, J = 6.7, H-14)/δC (14.2); one acetylene CH group at δH 2.00 (1H, s, H-1)/δC (65.6); three quaternary carbon groups at δC 68.2 (C-2), 67.0 (C-3), and 74.5 (C-4); and six methylenes at δH (1.30–2.64)/δC (22.7–33.6). The 1H and 13C-NMR spectroscopic data of compound 4 approached those of (6S,7S)-6,7-dihydroxytetradeca-1,3-diyne [36] except for an additional O-bearing methine in compound 4. The structure was further elucidated by the HMBC and 1H-1H COSY data (Figure 2). The 1H-1H COSY correlations between H-5/H-6/H-7/H-8/H-9, together with the HMBC correlations from H-8 to C-6, C-9 (δC 33.6), and C-10 (δC 26.0) and from H-6 to C-4 and C-5 (δC 24.6) suggested that compound 4 has three successive O-bearing CH groups located at C-6, C-7, and C-8. Thus, the structure of compound 4 was elucidated as tetradeca-1,3-diyne-6,7,8-triol.
Compound 5 was obtained as a yellow oil with a negative optical rotation [ α ] D 25 −16 (c 0.1, CHCl3). The molecular formula was determined as C22H30O8 by the HRESIMS peak at m/z 445.1859 [M + Na]+ (calcd for C22H30NaO8, 445.1833). The 1H-NMR spectrum (Table 2) showed four O-bearing methine groups at δH 5.11 (1H, q, J = 5.6 Hz, H-6), 5.39 (1H, dd, J = 4.7, 6.2 Hz, H-7), 5.18 (1H, dd, J = 4.7, 6.2 Hz, H-8), and 5.07 (1H, q, J = 6.4 Hz, H-9); one triplet methyl group at δH 0.86 (3H, t, J = 6.9 Hz, H-14); one acetylene CH group at δH 2.00 (1H, s, H-1); 10 methylene protons at δH 1.67–2.62; and 12 acetoxy methyl protons at δH 2.10–2.12. The 13C-NMR and DEPT spectra indicated the presence of 22 carbons—one terminal methyl carbon at δC 14.1 (C-14); four O-bearing methine carbons at δC 69.1 (C-6), 70.6 (C-7), 71.2 (C-8), and 71.5 (C-9); three quaternary acetylenic carbons at δC 68.0 (C-2), 67.6 (C-3), and 71.6 (C-4); one tertiary acetylenic carbon at δC 66.0 (C-1); and five methylene carbons at 21.8–31.5 ppm. All the assignments were supported by the HSQC experiments. A detailed comparison of the spectroscopic data of compounds 4 and 5 showed that in compound 5, there were four more acetyl groups [δC (170.0–170.6)/δH (2.10–2.12)] and one more O-bearing methine than in compound 4. The 1H-1H COSY correlations (Figure 2) of H-5/H-6/H-7/H-8 and H-9/H-10/H-11, as well as the HMBC correlations (Figure 2) from H-6/H-6″ to C-6′ (170.1), from H-7/H-7″ to C-7′ (170.0), from H-8/H-8″ to C-8′ (170.3), and from H-9/H-9″ to C-9 (170.6), indicated that compound 5 has four acetyls linked to C-6, C-7, C-8, and C-9, respectively. The other key correlations of HMBC and 1H-1H COSY are shown in Figure 2. Thus, the planar structure of compound 5 was elucidated as 6,7,8,9-tetraacetoxytetradeca-1,3-diyne.
Due to the high flexibly of the carbon chains, crystals were not obtained from compounds 15. Furthermore, the triple bonds contained compounds that were unstable and highly reactive [1], which resulted in all the isolates being chemically changed via oxidation and degradation. Therefore, the configurations of compounds 15 were not elucidated because of the materials’ instability and were reported as shown in Figure 1.
Compounds 17 were assessed for cytotoxic activity in the BEL-7402, K562, and SGC-7901 cancer cell lines, respectively. The results show that compounds 1 and 6 exhibited cytotoxicity in the above three human cancer lines, ranging from 14.3 to 45.4 μM (Table 3). Contrastingly, compound 4 displayed weak cytotoxicity in SGC-7901 and K562, with IC50 values of 26.2 ± 0.4 and 21.9 ± 0.3 μM, respectively. By comparison, the cytotoxicity in BEL-7401 and SGC-7901 was slightly enhanced when the double bond between C-9 and C-10 in compound 1 was oxidized to the epoxy group in compound 6, which could be due to the highly genotoxic effect of the epoxy group [37,38].

3. Materials and Methods

3.1. General Experimental Procedures

The 1H, 13C, and 2D NMR spectra were recorded on a Bruker AV III spectrometer (Bruker, Bremen, Germany) at either 500 MHz (1H) or 125 MHz (13C) using TMS as an internal standard. The HRMS were measured with an API QSTAR Pulsar mass spectrometer (Bruker). The UV spectra were performed on a Shimadzu UV-2550 spectrometer (Beckman, Brea, CA, USA). The IR absorptions were obtained on a Nicolet 380 FT-IR instrument (Thermo, Pittsburgh, PA, USA) using KBr pellets. The optical rotation was measured on a Rudolph Autopol III polarimeter (Rudolph, Hackettstown, NJ, USA). Silica gel (60–80, 200–300 mesh, Qingdao Marine Chemical Co. Ltd., Qingdao, China), ODS gel (20–45 μm, Fuji Silysia Chemical Co. Ltd., Durham, NC, USA), and Sephadex LH-20 (Merck, Darmstadt, Germany) were used for column chromatography. The TLC was conducted on pre-coated silica gel G plates (Qingdao Marine Chemical Co. Ltd.), and spots were detected by spraying with 10% H2SO4 in EtOH followed by heating.

3.2. Plant Material

The plant material was collected at the Chinese Academy of Tropical Agricultural Sciences, Haikou, China, in April 2014 and was identified as Swietenia macrophylla by Dr. Jun Wang (Institute of Tropical Bioscience and Biotechnology, Chinese Academy of Tropical Agricultural Sciences). The voucher specimen (No. DYTHXM201404) was deposited at the Institute of Tropical Bioscience and Biotechnology, Chinese Academy of Tropical Agricultural Science.

3.3. Extraction and Isolation

The roots of S. macrophylla (57.0 kg, dry weight) were first crushed and extracted with 95% EtOH (3 × 150.0 L) at room temperature and evaporated to yield EtOH extract (8.0 kg), which was then partitioned with H2O (40.0 L) and extracted with petroleum ether (PE) (40.0 L × 3) and EtOAc (40.0 L × 3), respectively. The EtOAc extract (3931.0 g) was subjected to silica gel (20 × 50 cm, 12.0 kg) vacuum liquid chromatography, and eluted with CHCl3/MeOH (v/v, 10:1; 40.0 L) to obtain fraction 1 (Fr.1). Silica gel (10 × 50 cm, 2.0 kg) vacuum liquid chromatography of Fr.1 (573.0 g) was eluted with PE/EtOAc (v/v, 1:0, 100:1, 50:1, 25:1, 10:1, 5:1, each 5.0 L, gradient) and CHCl3/MeOH (v/v, 1:0, 100:1, 50:1, 25:1, 10:1, 5:1, 0:1, each 5.0 L, gradient), respectively, which resulted in 17 sub-fractions (Fr.1.1−Fr.1.17). Fr.1.8 (19.5 g) was applied to an ODS gel (4.5 × 40 cm) and eluted with MeOH/H2O (v/v, 3:7, 2:3, 1:1, 3:2, 7:3, 4:1, 9:1, 1:0, each 5.0 L) to yield Fr.1.8.1−10. Fr.1.8.5 (2.9 g) was separated on a Sephadex LH-20 (3 × 100 cm) with CHCl3/MeOH as the eluent (v/v, 1:1; 1.5 L) to give Fr.1.8.5.1−5. Fr.1.8.5.1 (1.6 g) was separated on a silica gel column (3 × 45 cm, 160 g) and eluted with PE/EtOAc (v/v, 12:1) to yield compound 1 (30.0 mg) before it was eluted with a gradient of PE/acetone (v/v, 20:1, 15:1, 5:1, 1:1, 0.6 L of each) to give six fractions (Fr.1.8.5.1.1−6). Compound 7 (2.2 mg) from Fr.1.8.5.1.1 (32.0 mg) was obtained from a silica gel (1 × 12 cm, 6 g) eluting with PE/EtOAc (v/v, 8:1). Compound 6 (30.0 mg) from Fr.1.8.5.1.2 (380.0 mg) was obtained from a silica gel (2 × 30 cm, 60.0 g) eluting with PE/CHCl3 (v/v, 1:1). Fr. 1.8.5.1.3 (150.0 mg) was subjected to a silica gel (1 × 25 cm, 15.0 g) with PE/EtOAc (v/v, 5:1) to obtain compound 3 (8.0 mg). Fr. 1.8.5.1.4 (180.6 mg) was applied to a silica gel (1 × 30 cm, 18.0 g) eluting with PE/EtOAc (v/v, 4:1) to yield compound 2 (14.8 mg). Fr. 1.8.5.1.5 (121.4 mg) was purified by a silica gel (1 × 30 cm, 20.0 g) eluting with CHCl3/acetone (v/v, 5:1) to obtain compounds 5 (1.2 mg) and 4 (3.8 mg).
Heptadeca-9-ene-4,6-diyne-3,11-diol (1): yellow oil; UV (CH3OH) λmax (log ε) 280 (2.90), 270 (2.98) nm; IR (KBr) νmax 3435, 2924, 2234, 1639, 1384, 1020 cm−1; [ α ] D 25 +10 (c 0.1, CH3OH); 1H and 13C-NMR data: Table 1; HRESIMS m/z 285.1830 [M + Na]+ (calcd. C17H26NaO2 for 285.1831).
(E)-Heptadeca-8-ene-4,6-diyne-3,10,11-triol (2): yellow oil; UV (CHCl3) λmax (log ε) 286 (3.60), 270 (3.68) nm; IR (KBr) νmax 3433, 2927, 2250, 1645, 1382, 1026 cm−1; [ α ] D 25 −80 (c 0.1, CHCl3); 1H and 13C-NMR data: Table 1; HRESIMS m/z 301.1776 [M + Na]+ (calcd. C17H26NaO3 for 301.1780).
10-Methoxyheptadeca-4,6-diyne-3,9-diol (3): yellow oil; UV (CHCl3) λmax (log ε) 286 (2.77), 270 (2.78) nm; IR (KBr) νmax 3432, 2929, 2245, 1388, 1028 cm−1; [ α ] D 25 +80 (c 0.1, CHCl3); 1H and 13C-NMR data: Table 1; HRESIMS m/z 317.2089 [M + Na]+ (calcd. C18H30NaO3 for 317.2093).
Tetradeca-1,3-diyne-6,7,8-triol (4): yellow oil; UV (CHCl3) λmax (log ε) 298 (1.32), 240 (2.60) nm; IR (KBr) νmax 3434, 2926, 2241, 1384, 1018 cm−1; [ α ] D 25 −80 (c 0.1, CHCl3); 1H and 13C-NMR data: Table 2; HREIMS m/z 261.1476 [M + Na]+ (calcd for C14H22NaO3, 261.1467).
6,7,8,9-Tetraacetoxytetradeca-1,3-diyne (5): yellow oil; UV (CHCl3) λmax (log ε) 212 (4.44), 271 (3.55), 308 (3.12) nm; IR (KBr) νmax 2924, 2233, 1741, 1634, 1460, 1090 cm−1; [ α ] D 25 −16 (c 0.1, CHCl3); 1H and 13C-NMR data: Table 2; HREIMS m/z 445.1859 [M + Na]+ (calcd for C22H30NaO8, 445.1833).

3.4. Bioassay of Cytotoxic Activity

MTT assay, originally described by Mosmann [39], was used to quantitate the cytotoxicity of compounds 17. The human hepatocellular carcinoma cell line BEL-7402, human myeloid leukemia cell line K562, and human gastric carcinoma cell line SGC-7901, which were obtained from the cell bank of type culture collection of the Chinese Academy of Sciences, Shanghai Institute of Cell Biology, were cultured in RPMI 1640 medium supplemented with 10% fetal bovine serum at the conditions of 37 °C, 5% CO2, and 90% humidity. Paclitaxel was used as the positive control and DMSO was used as the negative control. Different concentrations of the test sample (each had triplicate wells) were designed as 0.1, 0.4, 1.6, 6.3, 25, and 100 μM. The logarithmic phase cells (90 μL) were selected to seed onto the 96-well plates at a concentration of 5 × 104 cell/mL. Then, 15 μL of MTT dissolved in PBS at 5 mg/mL was added to each well, and the system was incubated at 37 °C for 4 h. After that, the supernatant was discarded, and 100 μL of DMSO was added into each well. Finally, the OD value was measured by a MK3 Microtiter plate reader at a wavelength of 490 nm.

4. Conclusions

Seven polyacetylenes were isolated from the roots of S. macrophylla. Their structures were determined by spectroscopic analysis and comparing data in the literature. Furthermore, compounds 1 and 6 displayed weak cytotoxicity in the BEL-7402, SGC-7901, and K562 cell lines, and compound 4 showed a weak cytotoxic effect in the SGC-7901 and K562 cell lines.

Supplementary Materials

HRESIMS and NMR spectra for compounds 15 are available online.

Author Contributions

The list authors contributed to this work as follows: C.-N.M. processed the data, collected the plant samples, and prepared the manuscript. C.-N.M. and H.W. contributed to the structural elucidation. W.-L.M., H.-Q.C., and H.W. contributed to the revision of this manuscript. C.-H.C. and S.-P.L. conducted the bioassay experiments. The research was performed based on the planning of H.-F.D. and W.-L.M. All the authors approved the final version of the manuscript. Conceptualization, S.-P.L., W.-L.M., and H.-F.D.; Data curation, C.-N.M. and C.-H.C.; Formal analysis, H.W. and H.-Q.C.; Methodology, C.-N.M.; Project administration, W.-L.M. and H.-F.D.; Resources, W.-L.M. and H.-F.D.; Supervision, H.-Q.C., S.-P.L., W.-L.M., and H.-F.D.; Validation, H.W., H.-Q.C., and C.-H.C.; Original draft preparation, and C.-N.M.; and Review and editing of the manuscript, H.W.

Funding

This research was funded by the Innovative Research Team Grant of the Natural Science Foundation of Hainan Province (No. 2017CXTD020) and the Central Public-interest Scientific Institution Basal Research Fund for Chinese Academy of Tropical Agricultural Sciences (No. 17CXTD-15).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Konovalov, D.A. Polyacetylene compounds of plants of the Asteraceae, Family (Review). Pharm. Chem. J. 2014, 48, 613–631. [Google Scholar]
  2. Hansen, L.; Boll, P.M. Polyacetylenes in Araliaceae: Their chemistry, biosynthesis and biological significance. Phytochemistry 1986, 25, 285–293. [Google Scholar] [CrossRef]
  3. Hirakura, K.; Morita, M.; Nakajima, K.; Ikeya, Y.; Mitsuhashi, H. Three acetylated polyacetylenes from the roots of Panax ginseng. Phytochemistry 1991, 30, 3327–3333. [Google Scholar] [CrossRef]
  4. Hirakura, K.; Morita, M.; Nakajima, K.; Ikeya, Y.; Mitsuhashi, H. Three acetylenic compounds from roots of Panax ginseng. Phytochemistry 1992, 31, 899–903. [Google Scholar] [CrossRef]
  5. Yang, M.C.; Hakcheol, K.; Youngjoo, K.; Kangro, L.; Yang, H.O. Oploxynes A and B, polyacetylenes from the stems of Oplopanax elatus. J. Nat. Prod. 2010, 73, 801–805. [Google Scholar] [CrossRef] [PubMed]
  6. Chan, H.H.; Sun, H.D.; Reddy, M.V.; Wu, T.S. Potent alpha-glucosidase inhibitors from the roots of Panax japonicus C. A. Meyer var. major. Phytochemistry 2010, 71, 1360–1364. [Google Scholar] [CrossRef]
  7. Wang, C.Z.; Zhang, Z.Y.; Huang, W.H.; Du, G.J.; Wen, X.D.; Calway, T.; Yu, C.H.; Nass, R.; Zhao, J.; Du, W.; et al. Identification of potential anticancer compounds from Oplopanax horridus. Phytomedicine 2013, 20, 999–1006. [Google Scholar] [CrossRef]
  8. Chicca, A.; Pellati, F.; Adinolfi, B.; Matthias, A.; Massarelli, I.; Benvenuti, S.; Martinotti, E.; Bianucci, A.M.; Bone, K.; Lehmann, R.; et al. Cytotoxic activity of polyacetylenes and polyenes isolated from roots of Echinacea pallida. Brit. J. Pharmacol. 2008, 153, 879–885. [Google Scholar] [CrossRef]
  9. Quintana, N.; Weir, T.L.; Du, J.; Broeckling, C.D.; Rieder, J.P.; Stermitz, F.R.; Paschke, M.W.; Vivanco, J.M. Phytotoxic polyacetylenes from roots of Russian knapweed (Acroptilon repens (L.) DC.). Phytochemistry 2008, 69, 2572–2578. [Google Scholar] [CrossRef]
  10. Jung, H.J.; Min, B.S.; Park, J.Y.; Kim, Y.H.; Lee, H.K.; Bae, K.H. Gymnasterkoreaynes A-F, cytotoxic polyacetylenes from Gymnaster koraiensis. J. Nat. Prod. 2002, 65, 897–901. [Google Scholar] [CrossRef]
  11. Huang, H.Q.; Zhang, X.; Shen, Y.H.; Su, J.; Liu, X.H.; Tian, J.M.; Lin, S.; Shan, L.; Zhang, W.D. Polyacetylenes from Bupleurum longiradiatum. J. Nat. Prod. 2009, 72, 2153–2157. [Google Scholar]
  12. Bentley, R.K.; Jenkins, J.K.; Jones, E.R.H.; Thaller, V. Natural acetylenes. Part XXIX. Polyacetylenes from the Campanulaceae plant family. Tetrahydropyranyl polyacetylenic alcohols from the clustered bellflower (Campanula glomerata L.). J. Chem. Soc. C Organ. 1969, 5, 830–832. [Google Scholar] [CrossRef]
  13. Ishimaru, K.; Osabe, M.; Yan, L.; Fujioka, T.; Mihashi, K.; Tanaka, N. Polyacetylene glycosides from Pratia nummularia cultures. Phytochemistry 2003, 62, 643–646. [Google Scholar] [CrossRef]
  14. Koch, M.; Bugni, T.S.; Pond, C.D.; Sondossi, M.; Dindi, M.; Piskaut, P.; Ireland, C.M.; Barrows, L.R. Antimycobacterial activity of Exocarpos latifolius is due to exocarpic acid. Planta Med. 2009, 75, 1326–1330. [Google Scholar] [CrossRef]
  15. Bohlmann, F.; Rode, K.M. Polyacetylenverbindungen, 150. Notiz über die Polyineaus Pittosporum buchanani Hook. Fil. Eur. J. Inorg. Chem. 2010, 101, 1889–1891. [Google Scholar]
  16. Marles, R.J.; Farnsworth, N.R.; Neill, D.A. Isolation of a novel cytotoxic polyacetylene from a traditional anthelmintic medicinal plant, Minquartia guianensis. J. Nat. Prod. 1989, 52, 261–266. [Google Scholar] [CrossRef]
  17. Wakabayashi, N.; Spencer, S.L.; Waters, R.M.; Lusby, W.R. A Polyacetylene from Honduras Mahogany, Swietenia mahagoni. J. Nat. Prod. 1991, 54, 1419–1421. [Google Scholar] [CrossRef]
  18. Ning, J.; Di, Y.T.; Li, S.F.; Geng, Z.L.; He, H.P.; Wang, Y.H.; Wang, Y.Y.; Li, Y.; Li, S.L.; Hao, X.J. Polyynes from Toona ciliata var. ciliata and related cytotoxic activity. Helv. Chim. Acta 2011, 94, 376–381. [Google Scholar] [CrossRef]
  19. Zhang, B.; Wang, Y.; Yang, S.P.; Zhou, Y.; Wu, W.B.; Tang, W.; Zuo, J.P.; Li, Y.; Yue, J.M. Ivorenolide A, an unprecedented immunosuppressive macrolide from Khaya ivorensis: structural elucidation and bioinspired total synthesis. J. Am. Chem. Soc. 2012, 134, 20605–20608. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, Y.; Liu, Q.F.; Xue, J.J.; Zhou, Y.; Yu, H.C.; Yang, S.Y.; Zhang, B.; Zuo, J.P.; Li, Y.; Yue, J.M. Ivorenolide B, an immunosuppressive 17-membered macrolide from Khaya ivorensis: structural determination and total synthesis. Org. Lett. 2014, 16, 2062–2065. [Google Scholar] [CrossRef] [PubMed]
  21. Bernart, M.W.; Ii, J.H.C.; Balaschak, M.S.; Alexander, M.R.; Shoemaker, R.H.; Boyd, M.R. Cytotoxic falcarinol oxylipins from Dendropanax arboreus. J. Nat. Prod. 1996, 59, 748–753. [Google Scholar] [CrossRef]
  22. Koch, M.; Bugni, T.S.; Sondossi, M.; Ireland, C.M.; Barrows, L.R. Exocarpic acid inhibits mycolic acid biosynthesis in Mycobacterium tuberculosis. Planta Med. 2010, 76, 1678–1682. [Google Scholar] [CrossRef] [PubMed]
  23. Kobaisy, M.; Abramowski, Z.; Lermer, L.; Saxena, G.; Hancock, R.E.; Towers, G.H.; Doxsee, D.; Stokes, R.W. Antimycobacterial polyynes of Devil’s Club (Oplopanax horridus), a North American native medicinal plant. J. Nat. Prod. 1997, 60, 1210–1213. [Google Scholar] [CrossRef]
  24. Sara, I.; Yoel, K.; Shoshana, L.; Amnon, H.; Yossi, L. Petrosynol and petrosolic acid, two novel natural inhibitors of the reverse transcriptase of human immundeficiency virus from Petrosia sp. Tetrahedron 1993, 49, 10435–10438. [Google Scholar]
  25. Fusetani, N.; Sugano, M.; Matsunaga, S.; Hashimoto, K. H, K-ATPase inhibitors from the marine sponge siphonochalinatruncata: Absolute configuration of siphonodiol and two related metabolites. Tetrahedron Lett. 1987, 28, 4311–4312. [Google Scholar] [CrossRef]
  26. Pennington, T.D. Mahogany carving a future. Biologist 2002, 49, 204. [Google Scholar] [PubMed]
  27. Paiva, E.A. Anatomy, ultrastructure, and secretory activity of the floral nectaries in Swietenia macrophylla (Meliaceae). Am. J. Bot. 2012, 99, 1910–1917. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, S.K.; Chen, B.Y.; Li, H. Flora of China; Science Press: Beijing, China, 1997. [Google Scholar]
  29. Mulholland, D.A.; Parel, B.; Coombes, P.H. The chemistry of the Meliaceae and Ptaeroxylaceae of Southern and Eastern Africa and Madagascar. Curr. Org. Chem. 2000, 4, 1011–1054. [Google Scholar] [CrossRef]
  30. Tan, Q.G.; Luo, X.D. Meliaceous limonoids: Chemistry and biological activities. Chem. Rev. 2011, 111, 7437–7522. [Google Scholar] [CrossRef]
  31. Sun, Y.P.; Jin, W.F.; Wang, Y.Y.; Wang, G.; Morris-Natschke, S.L.; Liu, J.S.; Wang, G.K.; Lee, K.H. Chemical structures and biological activities of limonoids from the genus Swietenia (Meliaceae). Molecules 2018, 23, 1588. [Google Scholar] [CrossRef] [PubMed]
  32. Mi, C.N.; Mei, W.L.; Li, W.; Wang, J.; Cai, C.H.; Li, S.P.; Dai, H.F. Chemical constituents from the roots of Swietenia macrophylla King. J. Trop. Subtrop. Bot. 2017, 25, 610–616. [Google Scholar]
  33. Mi, C.N.; Li, W.; Chen, H.Q.; Wang, J.; Cai, C.H.; Li, S.P.; Mei, W.L.; Dai, H.F. Two new compounds from the roots of Swietenia macrophylla. J. Asian. Nat. Prod. Res. 2018. [Google Scholar] [CrossRef] [PubMed]
  34. Matsumori, N.; Kaneno, D.; Murata, M.; Nakamura, H.; Tachibana, K. Stereochemical determination of acyclic structures based on carbon-proton spin-coupling constants. A method of configuration analysis for natural products. J. Org. Chem. 1999, 64, 866–876. [Google Scholar] [CrossRef]
  35. Wu, Q.; Wu, C.M.; Long, H.L.; Chen, R.; Liu, D.; Proksch, P.; Guo, P.; Lin, W.H. Varioxiranols A−G and 19‑O‑methyl-22-methoxypre-shamixanthone, PKS and hybrid PKS-derived metabolites from a Sponge-associated Emericella variecolor Fungus. J. Nat. Prod. 2015, 78, 2461–2470. [Google Scholar] [CrossRef]
  36. Satoh, M.; Takeuchi, N.; Fujimoto, Y. Synthesis and absolute configuration of Panaxytriol. Chem. Pharm. Bull. 1997, 45, 1114–1116. [Google Scholar] [CrossRef]
  37. Lajovic, A.; Nagy, L.D.; Guengerich, F.P.; Bren, U. Carcinogenesis of urethane: simulation versus experiment. Chem. Res. Toxicol. 2015, 28, 691–701. [Google Scholar] [CrossRef]
  38. Klvana, M.; Bren, U. Aflatoxin B1–formamidopyrimidine DNA adducts: Relationships between structures, free energies, and melting temperatures. Molecules 2019, 24, 150. [Google Scholar] [CrossRef] [PubMed]
  39. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
Sample Availability: Not available.
Figure 1. Structures of compounds 17.
Figure 1. Structures of compounds 17.
Molecules 24 01291 g001
Figure 2. Key 1H-1H COSY and HMBC correlations of compounds 15.
Figure 2. Key 1H-1H COSY and HMBC correlations of compounds 15.
Molecules 24 01291 g002
Figure 3. Key ROESY interactions of compound 2.
Figure 3. Key ROESY interactions of compound 2.
Molecules 24 01291 g003
Table 1. 1H (500 MHz) and 13C-NMR (125 MHz) data for compounds 13 in CDCl3 (δ in ppm, J in Hz).
Table 1. 1H (500 MHz) and 13C-NMR (125 MHz) data for compounds 13 in CDCl3 (δ in ppm, J in Hz).
Position123
δHδCδHδCδHδC
10.99, t (7.4)9.5, CH31.01, t (7.4)9.5, CH31.01, t (7.4)9.5, CH3
21.72, m30.8, CH21.75, m30.8, CH21.73, m30.8, CH2
34.34, t (6.4)64.1, CH4.41, t (6.5)64.3, CH4.35, t (6.4)64.2, CH
4 77.3, C 83.5, C 77.0, C
5 69.7, C 69.5, C 69.8, C
6 65.0, C 74.4, C 66.3, C
7 78.6, C 76.7, C 77.8, C
83.09, dd (5.0, 10.2)18.3, CH25.86, d (15.9)110.5, CH2.51, dd (17.3, 6.1)
2.57, dd (17.3, 6.3)
24.8, CH2
95.50, overlapped135.3, CH6.30, dd (5.8, 15.9)146.3, CH3.70, td (6.3, 4.3)71.0, CH
105.48, overlapped124.4, CH4.00, td (5.8, 1.4)75.2, CH3.24, td (6.1, 4.3)82.1, CH
114.39 q (6.7)67.8, CH3.47, m74.4, CH1.55, m29.9, CH2
121.60, m; 1.44, m37.4, CH21.46, m33.2, CH21.28–1.29, m25.3, CH2
131.27–1.28, m25.3, CH21.27–1.28, m25.7, CH21.28–1.29, m29.4, CH2
141.27–1.28, m29.3, CH21.27–1.28, m29.4, CH21.28–1.29, m29.9, CH2
151.27–1.28, m31.9, CH21.27–1.28, m31.9, CH21.28–1.29, m31.9, CH2
161.27–1.28, m22.7, CH21.27–1.28, m22.7, CH21.28–1.29, m22.8, CH2
170.87 t (6.8)14.2, CH30.87 t (6.9)14.2, CH30.88 t (6.8)14.2, CH3
−OCH3 3.42, s58.5, CH3
Table 2. 1H (500 MHz) and 13C-NMR (125 MHz) data for compounds 4 and 5 in CDCl3 (δ in ppm, J in Hz).
Table 2. 1H (500 MHz) and 13C-NMR (125 MHz) data for compounds 4 and 5 in CDCl3 (δ in ppm, J in Hz).
Position45
δHδCδHδCPositionδHδC
12.00, s65.6, CH2.00, s66.0, CH6′ 170.1, C
2 68.2, C 68.0, C7′ 170.0, C
3 67.0, C 67.6, C8′ 170.3, C
4 74.5, C 71.6, C9′ 170.6, C
52.58, dd (17.4, 6.8)
2.64, dd (17.4, 6.5)
24.6, CH22.62, d (5.6)21.8, CH26″2.11 b, s20.8 a, CH3
64.13, br t (6.5)69.6, CH5.11, q (5.6)69.1, CH7″2.12 b, s21.0 a, CH3
73.49, m73.1, CH5.39, dd (4.7, 6.2)70.6, CH8″2.10 b, s20.9 a, CH3
83.82, m75.1, CH5.18, dd (4.7, 6.2)71.2, CH9″2.10 b, s20.8 a, CH3
91.55, m33.6, CH25.07, q (6.4)71.5, CH
101.33, m1.55, m26.0, CH21.55, m30.6, CH2
111.30, br s29.4, CH21.28, m24.6, CH2
121.30, br s31.9, CH21.28, m31.5, CH2
131.30, br s22.7, CH21.27, m22.5, CH2
140.89, t (6.7)14.2, CH30.86, t (6.9)14.1, CH3
a,b exchangeable.
Table 3. Cytotoxicity of compounds 17 in the human cancer cell lines.
Table 3. Cytotoxicity of compounds 17 in the human cancer cell lines.
CompoundCell Line, IC50 (μM)
BEL-7402SGC-7901K562
124.9 ± 0.345.4 ± 0.616.8 ± 0.1
4>5026.2 ± 0.421.9 ± 0.3
614.3 ± 0.433.4 ± 0.616.6 ± 0.4
2, 3, 5, 7>50>50>50
Paclitaxel a4.3 ± 0.14.3 ± 0.28.6 ± 0.1
a Positive control.

Share and Cite

MDPI and ACS Style

Mi, C.-N.; Wang, H.; Chen, H.-Q.; Cai, C.-H.; Li, S.-P.; Mei, W.-L.; Dai, H.-F. Polyacetylenes from the Roots of Swietenia macrophylla King. Molecules 2019, 24, 1291. https://doi.org/10.3390/molecules24071291

AMA Style

Mi C-N, Wang H, Chen H-Q, Cai C-H, Li S-P, Mei W-L, Dai H-F. Polyacetylenes from the Roots of Swietenia macrophylla King. Molecules. 2019; 24(7):1291. https://doi.org/10.3390/molecules24071291

Chicago/Turabian Style

Mi, Cheng-Neng, Hao Wang, Hui-Qin Chen, Cai-Hong Cai, Shao-Peng Li, Wen-Li Mei, and Hao-Fu Dai. 2019. "Polyacetylenes from the Roots of Swietenia macrophylla King" Molecules 24, no. 7: 1291. https://doi.org/10.3390/molecules24071291

Article Metrics

Back to TopTop